You are on page 1of 6

Ras activation by SOS: Allosteric regulation by altered fluctuation

dynamics
Lars Iversen et al.
Science 345, 50 (2014);
DOI: 10.1126/science.1250373

This copy is for your personal, non-commercial use only.

If you wish to distribute this article to others, you can order high-quality copies for your
colleagues, clients, or customers by clicking here.

Downloaded from www.sciencemag.org on July 5, 2014


Permission to republish or repurpose articles or portions of articles can be obtained by
following the guidelines here.

The following resources related to this article are available online at


www.sciencemag.org (this information is current as of July 5, 2014 ):

Updated information and services, including high-resolution figures, can be found in the online
version of this article at:
http://www.sciencemag.org/content/345/6192/50.full.html
Supporting Online Material can be found at:
http://www.sciencemag.org/content/suppl/2014/07/02/345.6192.50.DC1.html
This article cites 68 articles, 17 of which can be accessed free:
http://www.sciencemag.org/content/345/6192/50.full.html#ref-list-1
This article appears in the following subject collections:
Cell Biology
http://www.sciencemag.org/cgi/collection/cell_biol

Science (print ISSN 0036-8075; online ISSN 1095-9203) is published weekly, except the last week in December, by the
American Association for the Advancement of Science, 1200 New York Avenue NW, Washington, DC 20005. Copyright
2014 by the American Association for the Advancement of Science; all rights reserved. The title Science is a
registered trademark of AAAS.
R ES E A RC H | R E S EA R C H A R T I C LE S

MOLECULAR KINETICS (DH), and Pleckstrin homology (PH) domains


(10). Autoinhibition is thought to occur through
steric occlusion of the allosteric site, which can
Ras activation by SOS: Allosteric be released by interaction of the PH domain with
phosphatidylinositol-4,5-bisphosphate (PIP2) or
regulation by altered other negative lipids on the membrane (11, 12).
Several SOS mutations, including R552G in the

fluctuation dynamics helical linker, lead to weakened autoinhibition,


excessive Ras activation, and Noonan Syndrome
developmental disorders (13). (Single-letter ab-
Lars Iversen,1*† Hsiung-Lin Tu,1* Wan-Chen Lin,1 Sune M. Christensen,1† breviations for the amino acid residues are as
Steven M. Abel,2‡ Jeff Iwig,3 Hung-Jen Wu,1§ Jodi Gureasko,3‖ Christopher Rhodes,4 follows: A, Ala; C, Cys; D, Asp; E, Glu; F, Phe; G,
Rebecca S. Petit,1 Scott D. Hansen,1 Peter Thill,5 Cheng-Han Yu,6¶ Dimitrios Stamou,7 Gly; H, His; I, Ile; K, Lys; L, Leu; M, Met; N,
Arup K. Chakraborty,2,5,8,9,10,11 John Kuriyan,1,3,12 Jay T. Groves1,6,12,13# Asn; P, Pro; Q, Gln; R, Arg; S, Ser; T, Thr; V, Val;
W, Trp; and Y, Tyr. In the mutants, other amino
Activation of the small guanosine triphosphatase H-Ras by the exchange factor Son of acids were substituted at certain locations; for
Sevenless (SOS) is an important hub for signal transduction. Multiple layers of regulation, example, R552G indicates that arginine at po-
through protein and membrane interactions, govern activity of SOS. We characterized sition 552 was replaced by glycine.)
the specific activity of individual SOS molecules catalyzing nucleotide exchange in Most biological and biochemical studies of
H-Ras. Single-molecule kinetic traces revealed that SOS samples a broad distribution SOS activity have relied on bulk assays. Physi-
of turnover rates through stochastic fluctuations between distinct, long-lived (more cally distinct aspects of SOS regulation, such as
than 100 seconds), functional states. The expected allosteric activation of SOS by membrane recruitment and allosteric modula-
Ras–guanosine triphosphate (GTP) was conspicuously absent in the mean rate. tion of specific catalytic activity, are intrinsically
However, fluctuations into highly active states were modulated by Ras-GTP. This reveals convolved in such observations (supplementary
a mechanism in which functional output may be determined by the dynamical spectrum text S1). Furthermore, any stochastic variation
of rates sampled by a small number of enzymes, rather than the ensemble average. among SOS molecules, such as fluctuations be-

C
tween different activity states, is averaged out in
ellular membranes organize signal trans- Ras, Rho, and Arf GTPase subfamily GEFs is the the ensemble result. Because many signaling pro-
duction, serving as platforms for protein release of autoinhibition mediated by GTPase cesses in cells involve small numbers of mole-
interactions as well as direct modulators and membrane binding. The molecular mech- cules, the ability to average over large ensembles
of enzymatic function (1, 2). The activa- anisms underlying these regulatory couplings is not a benefit that live cell-signaling networks
tion of lipid-anchored guanosine triphos- remain poorly understood, in large part because necessarily enjoy (14–17). Stochastic variation it-
phatases (GTPases) of the Ras superfamily by of the intrinsic experimental challenge of work- self, rather than ensemble average properties, can
cytosolic guanine nucleotide exchange factors ing within a membrane environment. We recon- be a viable mechanism of regulation (18). To gain
(GEFs) represents a broadly important class of stituted the inner-leaflet signaling geometry of a clear understanding of SOS activity and its
membrane-localized signaling reactions. Con- Son of Sevenless (SOS)–catalyzed nucleotide ex- regulation on membranes will require direct ob-
trol of GEF activity is multilayered and involves change in H-Ras on supported membranes that servations of individual SOS molecules function-
membrane recruitment, lateral interactions on were partitioned into arrays of two-dimensional ing in a membrane environment.
the membrane surface, as well as allosteric reg- corrals by lithographically defined chromium dif-
ulation (3, 4). A recurring feature across several fusion barriers. With this system, we monitored Single-molecule SOS activity assay
the real-time catalytic activity of individual SOS We developed an assay platform to observe real-
1
Howard Hughes Medical Institute, Department of Chemistry, molecules in order to observe the functional time single-molecule activity of SOS with Ras
University of California, Berkeley, Berkeley, CA 94720, USA. mechanisms of allosteric activation and auto- in a partitioned, supported membrane (Fig. 1B).
2
Department of Chemical Engineering, Massachusetts inhibition on the membrane surface. H-Ras(C118S, 1 to 181) (referred to as Ras from
Institute of Technology (MIT), Cambridge, MA 02139, USA.
3
Howard Hughes Medical Institute, Department of Molecular
SOS is widely distributed in mammalian cells. here on) was linked to the membrane through
and Cell Biology, University of California, Berkeley, Berkeley, In vivo, inactive cytosolic SOS is recruited to the maleimide coupling of Cys181 to supported lipid
CA 94720, USA. 4Department of Mechanical Engineering, plasma membrane in response to ligand bind- bilayers (SLBs), resulting in stably bound, lat-
University of California, Berkeley, Berkeley, CA 94720, USA. ing by receptors on the cell surface. There, it erally mobile Ras that is fully functional with
5
Department of Chemistry, MIT, Cambridge, MA 02139, USA.
6
Mechanobiology Institute, National University of Singapore,
activates membrane-tethered Ras by catalyzing respect to SOS activation (19, 20). We also used
Singapore. 7Department of Chemistry and Nano-Science the exchange of Ras-bound guanosine diphos- an H-Ras construct truncated at Cys184 [H-Ras
Center, University of Copenhagen, Copenhagen, Denmark. phate (GDP) with guanosine triphosphate (GTP), (C118S, 1 to 184)], which contains both native
8
Department of Biological Engineering, MIT, Cambridge, MA which triggers the mitogen-activated protein ki- cysteine palmitoylation sites (181 and 184) in the
02139, USA. 9Ragon Institute of Massachusetts General
Hospital, MIT, and Harvard, Cambridge, MA 02139, USA.
nase (MAPK) cascade (5). SOS is activated by Ras- hypervariable region (HVR), which we linked to
10
Department of Physics, MIT, Cambridge, MA 02139, USA. GTP binding to an allosteric site, located between lipids through maleimide chemistry. Membranes
11
Institute for Medical Engineering and Science, MIT, the Cdc25 and Ras exchanger motif (REM) domains were composed primarily of L-a-phosphatidylcholine
Cambridge, MA 02139, USA. 12Physical Biosciences and in the catalytic core termed SOScat (C) (Fig. 1A) (Egg, Chicken) (Egg-PC) with 2 to 3% anionic
Materials Sciences Divisions, Lawrence Berkeley National
Laboratory, Berkeley, CA 94720, USA. 13Berkeley Education
(6). This allosteric activation depends sensitively 1,2-dioleoyl-sn-glycero-3-phospho-L-serine lipids
Alliance for Research in Singapore, 1 Create Way, CREATE on the nucleotide state of Ras (7) and contributes (DOPS) and in some cases PIP2, which bind the
tower level 11, University Town, Singapore 138602. an important aspect of SOS biology. Function- PH domain on SOS and release autoinhibition
*These authors contributed equally to this work and are listed in ally, this is thought to enable SOS to operate as (19). Molecules within the supported membrane
alphabetical order. †Present address: Department of Chemistry,
University of Copenhagen, Copenhagen, Denmark. ‡Present
an analog-to-digital converter through a Ras-GTP were confined in micrometer-scale corrals by lith-
address: Department of Chemical and Biomolecular Engineering, positive-feedback loop operating at the mem- ographically defined barriers to lateral mobility,
University of Tennessee, Knoxville, TN 37996, USA. §Present brane, such as during T cell activation (8, 9). prefabricated onto the underlying substrate (21).
address: Department of Chemical Engineering, Texas A&M However, the nucleotide specificity of allosteric The barriers trapped membrane-tethered Ras
University, College Station, TX 77843, USA. ||Present address:
Epizyme, 400 Technology Square, Cambridge, MA 02139, USA.
activation of the catalytic site remains poorly and SOS (through its interactions with Ras and
¶Present address: Department of Anatomy, The University of Hong understood. SOS activation is autoinhibited by membrane lipids) within individual corrals. In
Kong, Hong Kong. #Corresponding author. E-mail: jtgroves@lbl.gov its N-terminal Histone fold (H), Dbl homology each corral, the membrane remained entirely

50 4 JULY 2014 • VOL 345 ISSUE 6192 sciencemag.org SCIENCE


R ES E A RC H | R ES E A R C H A R T I C L E S

fluid, thus permitting SOS and Ras to diffuse surement of Ras lateral mobility with FCS and SOS binds to Ras on the membrane surface and,
freely while SOS processively catalyzed nucleo- fluorescence recovery after photobleaching (FRAP) in the absence of free nucleotide, becomes trapped
tide exchange. Ras surface density was calibrated (fig. S2) ensured that fluid bilayers were parti- (3). After free SOS was rinsed from the system,
with fluorescence correlation spectroscopy (FCS) tioned by leak-free corrals. unlabeled nucleotide was flowed in, and the ex-
(22) and subsequently measured in each corral SOS was introduced into the system from change reaction commenced. By loading Ras with
by means of epifluorescence imaging (fig. S1). solution in a transient pulse traveling through Atto488-labeled fluorescent nucleotide 2′/3′-O-(2-
Ras surface densities in these experiments ranged the flow cell. We examined several SOS con- aminoethyl-carbamoyl)-guanosine-5′-diphosphate
from several hundred to 1000 Ras molecules per structs derived from a truncated SOS containing or -[(b,g)-imido]triphosphate (Atto488-EDA-GDP
square micrometer, corresponding to the broad the H, DH, PH, and C domains but lacking the or Atto488-EDA-GppNp, respectively; referred
Ras density range measured in vivo (8, 19). Mea- C-terminal Grb2 binding domain (SOS-HDPC). to as GDP-488 and GTP-488, respectively, from
here on), catalytic exchange with nonfluorescent
nucleotide from solution (GDP or GTP) could be
directly observed as a local, corral-confined de-
crease of surface fluorescence. Constant flow during
the exchange reaction ensured removal of unbound
fluorescent nucleotides, allowing the reaction ki-
netics to be recorded by means of wide-field im-
aging of the fluorescence decay. Fluorescent and
nonhydrolysable nucleotide analogs did not sub-
stantially perturb SOS activity (20).
Representative images taken during SOScat-
catalyzed turnover showed a small percentage
of active corrals (those with a SOS molecule)
undergoing nucleotide exchange. These develop
a clear negative contrast relative to inactive cor-
rals (those without a SOS molecule) (Fig. 1C and
movie S1). In all experiments, SOS concentrations
were adjusted so that >95% of active corrals con-
tained exactly one enzyme (fig. S3) (20). Total in-
ternal reflection fluorescence microscopy (TIRFM)
of Atto647N-labeled SOScat (SOScat-647) revealed
a clear correspondence between active dark cor-
rals and individually confined SOScat enzymes
(Fig. 1D and movie S2), which were laterally mob-
ile within corrals (fig. S4) (20). However, it was
not necessary to label SOS to run the assay. By
imaging arrays of corrals on the surface, the ex-
change reactions of hundreds of individually con-
fined enzymes were monitored in parallel (Fig.
1E). Quantitative analysis of fluorescence decays
(fig. S5) (20) allowed individual kinetic traces of
SOS activity to be analyzed (Fig. 1F). The total
number of molecules tracked in an experimental
run was similar to that in some bulk experiments
Fig. 1. Platform for single-enzyme kinetics. (A) Schematic showing the crystal structure and domain (19), except that the individual contribution of
architecture of SOS-HDPC (surface and illustrated rendering) with Ras molecules (yellow, illustrated every single molecule was observed.
rendering) modeled onto the allosteric and catalytic sites, facing a lipid bilayer. SOS-HDPC is de- SOScat enzymes were highly processive on the
picted in a sterically closed and auto-inhibited conformation. The actual on-membrane configura- membrane surface. Hundreds to several thousand
tion is expected to be more open. The SOS-HDPC crystal structure, the HVR region of Ras, and the Ras could be activated by one SOScat molecule
1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) bilayer structure are adapted from pub- during a single membrane residency period (Fig.
lished works (20). (B) Scheme of the experimental setup. Nanofabricated chromium metal lines 1G). Comparable results for SOS activity were
(10 nm high and 100 nm wide) partition a supported bilayer into micrometer-scale corrals, each observed with both the single- (C118S, 1 to 181)
containing lipid-anchored Ras loaded with fluorescent nucleotide at densities from hundreds to and double-anchored (C118S, 1 to 184) Ras con-
approximately a thousand molecules per square micrometer. When a single SOS engages Ras at structs (fig. S6). Because SOScat lacks any of the
the allosteric site, the catalytic site is free to turn over the remaining Ras in the corral, replacing other membrane binding domains, this confirms
fluorescent with nonfluorescent nucleotide and leading to a confined decrease of emission in- that Ras alone is sufficient to stably recruit SOS
tensity. (C) Wide-field epifluorescence image of fluorescently loaded Ras before injection of SOS to the membrane. A W729E point mutation in
(left) shows no initial dark corrals. Injection of a pulse of SOS followed by continuous flow of SOScat that abolishes Ras binding in the allo-
nonfluorescent nucleotide (right) leads to enzymatic turnover in a subset of corrals. (D) False color steric pocket (10, 19) inhibited processive activity
overlay of fluorescently loaded Ras emission (red) and TIRFM image of fluorescently labeled (fig. S7). Bivalent interaction with Ras through
SOScat-Atto647N (green) reveals colocalization of dark corrals and single SOScat enzymes. (E) Zoomed- both the catalytic and allosteric sites was required
out view showing a field of 1 × 1 mm2 corrals with SOS activity. (F) Collection of single-corral kinetic for sustained membrane localization of SOScat.
traces from an array of SOScat turning over Ras-GTP, showing the percentage of fluorescent Ras in the
corral as a function of time (blue traces). The black traces represent signals from empty corrals Dynamic heterogeneity of SOS activity
without enzymatic activity. (G) Histogram of total Ras-GTP-488 turnovers by individual SOScat Kinetic traces of nucleotide exchange from in-
enzymes. Scale bars, 10 mm (C) and (D); and 20 mm (E). Lipid composition (in molar percent): Egg-PC/ dividual SOS molecules revealed discrete tran-
MCC-PE/DOPS/TR-DHPE = 93.99/3/3/0.01. sitions between well-defined catalytic states.

SCIENCE sciencemag.org 4 JULY 2014 • VOL 345 ISSUE 6192 51


R ES E A RC H | R E S EA R C H A R T I C LE S

Traces from two representative SOS molecules, or PIP2 in these experiments) (10, 19, 28). How- steric site is surprisingly weak over most of the
each in its own corral, are plotted in Fig. 2A (more ever, we observed that the N-terminal domains range of the observed distribution of rates. Rate
traces are provided in fig. S8). One molecule ran still influence SOS activity by allosterically sup- histograms for SOScat, SOS-HDPC, and SOS-
for more than 1200 s with a single catalytic rate pressing fluctuations that populate multiple HDPC(R552G) are shown in Fig. 3, A, B, and C,
(1.1 molecules/s). The other molecule shown high-activity states, which is consistent with a respectively. Although small differences in the
started out with a similar rate, abruptly changed mechanism of conformational selection (supple- distributions and most probable rates do exist,
to 4 molecules/s at ~400 s, then abruptly changed mentary text S2) (29–31). This represents a pre- it is unlikely that these minor effects would be
back after another ~500 s. Using a change point viously unknown layer of SOS regulation and consequential in the context of a living cell. Some
algorithm (20, 23) to detect transitions revealed emphasizes that not just the allosteric site, but cellular signaling processes can be triggered by
such events in 30% of traces for SOScat turning also the N-terminal domains, communicate di- as little as a few individual ligand-receptor inter-
over Ras-GDP (fig. S9). Changes in turnover rate rectly with the catalytic site. actions (15–17). In such cases, only a small num-
were not associated with changes in lateral dif- The Noonan syndrome associated R552G SOS ber of SOS molecules are likely to be involved as
fusion of SOS (fig. S10). Because the transitions mutation leads to excessive activation of Ras well. Variation of the ensemble mean (stochastic
were discrete and the individual states exhib- in vivo (13) and increases the apparent rate of noise) becomes large when sample sizes become
ited well-defined kinetic rates, we suggest that catalysis in bulk assays, but only when membranes small, and thus, small differences in the mean
these are the result of transitions between dif- are present (19). In vesicle assays, Ras activation is rates become obscured in stochastic noise (sup-
ferent stable configurations of the protein com- monitored as a function of the total amount of plementary text S4 and fig. S13).
plex itself (SOScat with Ras bound in its allosteric SOS in the system, but the fraction of SOS actually However, the dynamics of SOS fluctuations
site). Such conformational fluctuations have been recruited to the membrane is unknown. The single- do exhibit differences, depending on whether
traditionally observed on the microsecond to mil- molecule rate histogram for SOS-HDPC(R552G) Ras-GTP or Ras-GDP is bound in the allosteric
lisecond time scale (24–26), but longer-lived dy- shows that the R552G mutation is not activating site. Whereas SOS accesses the rare, highly active
namic conformers exist in proteins (27), extending at the level of SOS specific activity (Fig. 2C) and states in either case, individual molecules linger
well into the minute time scale (supplementary results in a similar low-fluctuation frequency as for longer periods in these active states when in-
text S2). seen for SOS-HDPC (fig. S9). Taken together with teracting allosterically with Ras-GTP. This is made
By identifying individual functional substates, this observation, the accelerated turnover of Ras- vivid if we examine the total Ras activation by
we constructed the lifetime-weighted probabil- GTP by SOS-HDPC(R552G) reported from vesicle individual SOS molecules. In Fig. 3, D, E, and F,
ity histogram of distinct catalytic rates sampled assays (19) is likely to be a result of increased each observed catalytic state is plotted on the
by hundreds of SOScat enzymes (Fig. 2B) (20). surface binding from solution. Abnormally high basis of its specific activity and the Ras surface
The histogram spans almost 2 orders of magni- membrane recruitment of SOS-HDPC(R552G) may density on which it was observed. The total Ras
tude, with a broad peak at ~1 s−1 and a wide contribute to the pathogenic hyperactivation of turnover, equivalent to the number of Ras mol-
shoulder and tail of sparsely populated states Ras signaling in vivo as well. ecules activated by this particular SOS molecule
extending toward higher rates. The histogram while in this state, is represented by the size
of apparent rates from inactive corrals, arising Allosteric regulation of SOS of the mark. Ras-GTP allosterically promotes
from intrinsic nucleotide release and photo- A second key observation is that the expected high-activity long-lived states at the expense
bleaching, represents the limit of resolution in activation of SOS by Ras-GTP bound in its allo- of low-activity shorter-lived states (figs. S14
this experiment (20). Even in a single catalytic
state, stochastic variation resulting from the
discrete nucleotide exchange events will lead to
a real distribution of turnover rates over time 0.0 1.0 SOScat
No enzyme
ln(normalized int.)

and between enzymes. We estimated this in-


Relative probability

Stochastic variation
trinsic stochastic noise through simulations of a
one-state system (supplementary text S3 and
fig. S11). The overall distribution of SOScat turn- -0.5 0.5
over rates is much broader than either the ex-
perimental or intrinsic stochastic noise limits.
However, it is likely that the assay does not re-
solve individual state peaks, giving instead the -1.0 0.0
appearance of a broad continuous distribution. 0 500 1000 1500 0 2 4 6 8

SOS regulation through Time (s) Rate (molecules/s)


intradomain interaction Fig. 2. Allosteric autoinhibition by N-terminal do-
0.4 SOScat
A key finding comes from a comparison of rate mains of SOS. (A) Logarithmic plots of single-enzyme
SOS-HDPC
histograms for SOScat and SOS-HDPC (Fig. 2C SOS-HDPC(R552G) kinetic traces. Discrete activity levels of SOScat are
0.3
and fig. S12). The N-terminal H, DH, and PH identified as linear segments. Some SOScat enzymes
Probability

domains have a pronounced effect on the dis- remain in a single functional state throughout the
tribution of rates sampled by SOS. Whereas the 0.2
reaction (straight trace, green), whereas others fluctu-
peak of the rate distribution for SOS-HDPC ate between different states (kinked trace, blue). The
overlaps with that of SOScat, the extended tail 0.1 black trace (square) represents corral that does not
of molecules with faster rates is strongly sup- undergo SOScat-catalyzed turnovers. (B) Normalized
pressed, resulting in a sharper histogram. The 0.0 and lifetime-weighted probability distributions of cata-
N-terminal domains also dampen the frequency 0 2 4 6 8 lytic rates sampled by (i) the SOScat ensemble (black),
with which state transitions occur, dropping to Rate (molecules/s) (ii) corrals without enzyme (brown), and (iii) modeled
10% of kinetic traces with transitions for SOS- single-state enzyme stochastic variation (orange).
HDPC (fig. S9). These observations were made (C) Catalytic rate probability distributions from SOScat (black), SOS-HDPC (green), and SOS-HDPC
under conditions in which autoinhibition was (R552G) (purple). The N-terminal domains auto-inhibit SOS specific activity. A Noonan syndrome–
previously thought to be released by interactions associated point mutation in SOS-HDPC(R552G) (purple) does not relieve inhibition. Lipid composition
between the PH domain and anionic lipids (PS (in molar percent): Egg-PC/MCC-PE/DOPS/TR-DHPE = 93.99/3/3/0.01.

52 4 JULY 2014 • VOL 345 ISSUE 6192 sciencemag.org SCIENCE


R ES E A RC H | R ES E A R C H A R T I C L E S

and S15). In cells, these highly active individ- regulation, we conducted a series of stochastic brane signaling cluster (34)] were chosen from
ual molecules would produce all of the Ras they simulations, beginning with a simple enzymatic the same distribution, but their rates of switch-
activate in the same location (supplementary system that does not account for feedback reg- ing between states were different. Histograms of
text S5) and over a short time. This is especially ulation (Fig. 4, A and B). The simple system simulation results for many such reaction groups
consequential in a signaling context, in which demonstrates the functional importance of rate showed that for small numbers of enzymes (ana-
GTPase-activating proteins (GAPs) catalyze the fluctuations and the existence of long-lived active log of weak stimulation), slower switching rates
hydrolysis of GTP to GDP, thus inactivating Ras states. We consider a substrate, S, that is converted can enable some groups to become activated,
at a basal level. The effects of such GAP activity to its active state, S*, by an enzyme E, which can whereas activity at the average rate would not
would render small bursts of Ras activation sample different possible states of activity (sup- (Fig. 4B).
inconsequential but may not keep pace with the plementary text S6 and fig. S16). A deactivating We also investigated a coarse-grained mod-
highly active molecules. In such a situation, the enzyme (analog of RasGAP) can convert S* to S. el of lymphocyte SOS signaling, which has been
average activity of the entire SOS ensemble is We first carried out Gillespie simulations (33) used to interpret bimodal early signaling events
relatively unimportant because the rare, highly with single enzymes, with different trajectories in lymphocytes (supplementary text S7) (8). This
active individuals dominate the overall system corresponding to different choices of the en- model includes feedback regulation through
behavior. Furthermore, the important functional zyme activity and switching rates. In one set of Ras-GTP and the interplay between SOS and the
consequences of feedback regulation of SOS by trajectories, the enzymes were allowed to sample RasGEF Ras guanyl nucleotide-releasing protein
Ras-GTP observed in computer simulations and only low-activity states (allosteric Ras-GDP–bound (RasGRP), as well as the signal attenuation through
cellular experiments (8) could well be mediated SOS), whereas in another set, an additional high- GAP activity (8). The resulting histograms of cel-
by the long-lived active states that allosteric reg- ly active state was sampled (allosteric Ras-GTP); lular Ras-GTP levels are shown in Fig. 4C, each
ulation by Ras-GTP enables but that by Ras-GDP in another set of simulations, the enzymes sam- containing thousands of simulated cells. The ex-
does not. Ras reportedly forms dimers on mem- pled identical enzymatic rate distributions but istence of stochastic activity fluctuations between
brane surfaces (32). Such dimers could poten- transitioned between catalytic states at different long-lived states in the simulated SOS-like GEF
tially affect SOS activity and, depending on the rates. Our simulations demonstrate that placing shifted the regime of bimodal Ras-GTP signaling
two-dimensional dimerization affinity on mem- additional weight in the highly active tail of the (gray-shaded histograms) and broadened the
brane surfaces, could contribute to the mild distribution or having enzymes with long-lived ability of the network to exhibit bistability. This
Ras density–dependence of SOS activity seen states can lead to fast threshold-crossing by a effect propagated through the network as a whole.
in Fig. 3D. subset of the enzymes, whereas a collection of Bimodality resulting from static SOS (Fig. 4D,
enzymes all operating at an increased but aver- red lines) or fluctuating SOS (Fig. 4D, blue lines)
Effect of fluctuation dynamics age catalytic rate are unable to support activation differed across the SOS and RasGRP concentra-
on the signaling network (Fig. 4A). We also made simulations in which tion matrix (Fig. 4D). Gray-shaded histograms
To explore the general feasibility of rate fluc- the rates of a number of enzymes in a single re- highlight regions of the parameter space where
tuations that provide a mechanism of allosteric action group [such as a cell or a localized mem- only fluctuating SOS enzymes gave rise to bi-
modal Ras-GTP distributions, illustrating how
the threshold of the overall signaling network
is fine-tuned by stochastic rate fluctuations of
GDP -> GDP
10-1
GTP -> GTP 10-1 10-1 a single constituent enzyme—in this case, SOS.
Probability
Probability

Probability

10-2 Various models have been advanced to explain


10-2 10-2
10-3 how Ras-GTP (35, 36) and other cellular signals
10-4
10-3 10-3 (37, 38) are spatiotemporally shaped to process
10-5 10-4 10-4 information, for example, for analog or digital
0 5 10 15 20 0 5 10 0 5 10
response to stimuli. Stochastic rate fluctuations,
Rate (molecules/s) Rate (molecules/s) Rate (molecules/s) as reported here for SOS, constitute another
mechanism for tuning cellular signaling dynam-
20 10 10
ics (39).
Dynamic heterogeneity has been reported for
multiple enzymes (40–42). However, the dura-
Rate (molecules/s)

5
Rate (molecules/s)

10 5
Rate (molecules/s)

tion of dynamic SOS states we observed [several


minutes (Fig. 3B)] is longer than that reported
0 0 0 for other enzyme systems [up to tens of seconds
20 10 10
(41, 42)]. The SOScat fluctuation behavior shows
n n
n that dynamic conformational modes in signal-
12500 3400
2820
10 5 5 ing enzymes may extend well into time scales
7500 2040 1692 comparable with receptor signaling processes
2500 680
(37, 39). Evolutionary pressure is required to evolve
564
30 7 6
0 0 0
400 800 1200 400 800 1200 and maintain increased enzymatic efficiency (43),
400 800 1200
Ras density
so high-activity states are likely to be biologically
Ras density Ras density
(molecules/µm2) (molecules/µm2) (molecules/µm2) important. SOS is the key enzyme involved
in bimodal switching of Ras-GTP levels during
Fig. 3. Nucleotide specificity of allosteric activation of SOS. (A to C) The nucleotide-dependent rate T cell receptor–triggering in T cells (8, 44).
distributions for (A) SOScat, (B) SOS-HDPC, and (C) SOS-HDPC(R552G) on a logarithmic scale. For all It is certainly possible that rare high-activity
constructs, Ras-GTP bound in the allosteric site of SOS (blue traces) has a small activating effect relative states could function in threshold-crossing and
to Ras-GDP in the allosteric site (red traces). (D to F) Scatterplots of SOS turnover rate as a function of bistability, leading to robust Ras-GTP positive-
Ras surface density for (D) SOScat, (E) SOS-HDPC, and (F) SOS-HDPC(R552G). Each point in the scat- feedback activation of SOS and robust re-
terplots represents an observed catalytic state, and the area of the point represents the number of Ras sponse to antigen.
turned over by that catalytic state. For all constructs, Ras-GTP bound in the allosteric site of SOS (blue More fundamentally, we developed a single-
points) results in longer-lived high-activity states capable of highly processive catalysis. Lipid composition molecule enzymatic assay that enables detailed
(in molar percent): Egg-PC/MCC-PE/DOPS/TR-DHPE = 93.99/3/3/0.01. observation of the activation of membrane-linked

SCIENCE sciencemag.org 4 JULY 2014 • VOL 345 ISSUE 6192 53


R ES E A RC H | R E S EA R C H A R T I C LE S

16. B. N. Manz, B. L. Jackson, R. S. Petit, M. L. Dustin,


J. Groves, Proc. Natl. Acad. Sci. U.S.A. 108, 9089–9094
(2011).
17. G. P. O’Donoghue, R. M. Pielak, A. A. Smoligovets, J. J. Lin,
J. T. Groves, eLife 2, e00778 (2013).
18. P. J. Choi, L. Cai, K. Frieda, X. S. Xie, Science 322,
442–446 (2008).
19. J. Gureasko et al., Nat. Struct. Mol. Biol. 15, 452–461
(2008).
20. Materials and methods are available as supplementary
materials on Science Online.
21. J. T. Groves, N. Ulman, S. G. Boxer, Science 275, 651–653
(1997).
22. J. T. Groves, R. Parthasarathy, M. B. Forstner, Annu. Rev.
Biomed. Eng. 10, 311–338 (2008).
23. D. L. Ensign, V. S. Pande, J. Phys. Chem. B 114, 280–292
(2010).
24. D. D. Boehr, R. Nussinov, P. E. Wright, Nat. Chem. Biol. 5,
789–796 (2009).
25. K. Henzler-Wildman, D. Kern, Nature 450, 964–972
(2007).
26. K. A. Henzler-Wildman et al., Nature 450, 913–916
(2007).
27. W. Min, G. Luo, B. J. Cherayil, S. C. Kou, X. S. Xie, Phys. Rev.
Lett. 94, 198302 (2005).
28. C. Zhao, G. Du, K. Skowronek, M. A. Frohman, D. Bar-Sagi,
Nat. Cell Biol. 9, 706–712 (2007).
29. A. del Sol, C. J. Tsai, B. Ma, R. Nussinov, Structure 17,
1042–1050 (2009).
30. J. P. Changeux, Annu. Rev. Biophys. 41, 103–133
(2012).
Fig. 4. Modeling effects of stochastic transitions between long-lived enzymatic states by using a 31. P. Csermely, R. Palotai, R. Nussinov, Trends Biochem. Sci.
minimal signaling model or broader network model. (A) Fraction of trajectories (f) reaching a 35, 539–546 (2010).
32. J. Güldenhaupt et al., Biophys. J. 103, 1585–1593
threshold amount (NS*) of product S* during activation by single GEF-like enzymes fluctuating with
(2012).
switching frequencies (gS). GAP-like enzymes simultaneously deactivate S* to S at a fixed rate. The 33. D. T. Gillespie, J. Comput. Phys. 22, 403–434 (1976).
ensembles of enzymes represented by the three traces have identical mean activity levels and differ only 34. N. C. Hartman, J. T. Groves, Curr. Opin. Cell Biol. 23, 370–376
in the rate with which enzymes fluctuate between activity states (details of the model are provided in (2011).
35. A. S. Harding, J. F. Hancock, Trends Cell Biol. 18, 364–371
supplementary text S6). Enzymes with a fast switching rate (green trace) do not reach the signaling
(2008).
threshold, whereas enzymes with a lower switching rate (orange trace) do. Enzymes that do not switch 36. B. N. Kholodenko, J. F. Hancock, W. Kolch, Nat. Rev. Mol. Cell
between states on the time scale of the simulation show the most rapid activation and threshold Biol. 11, 414–426 (2010).
crossing. Additionally, no threshold crossing is observed for static enzymes operating at the average 37. H. Wu, Cell 153, 287–292 (2013).
38. H. E. Grecco, M. Schmick, P. I. Bastiaens, Cell 144, 897–909
catalytic rate. (B) Probability distributions of (NS*) for various ensemble sizes (Nenzyme = 4, 16, or 64),
(2011).
with slow switching (brown trace, gS = 10−3), fast switching (green trace, gS = 10−1), or enzymes 39. J. E. Purvis, G. Lahav, Cell 152, 945–956 (2013).
operating only at the average catalytic rate (blue trace). In all cases, the average catalytic rate is the 40. H. P. Lu, L. Xun, X. S. Xie, Science 282, 1877–1882
same. At low (4) and intermediate (16) copy number, the activation pattern depends strongly on the (1998).
switching rate. At low copy number, slow switching gives rise to a population of highly activating 41. B. P. English et al., Nat. Chem. Biol. 2, 87–94 (2006).
42. M. K. Prakash, R. A. Marcus, Proc. Natl. Acad. Sci. U.S.A. 104,
enzymes (resulting in more than 400 S*). (C) Distributions of Ras-GTP resulting from coarse-grained 15982–15987 (2007).
modeling of the Ras-SOS signaling network (details of the model are provided in supplementary text S7). 43. W. J. Albery, J. R. Knowles, Biochemistry 15, 5631–5640
Each histogram contains 8000 modeled cells. Left column shows simulations without catalytic rate (1976).
fluctuations, gS = 0. Middle column shows simulations with catalytic rate fluctuations for which the 44. J. P. Roose, M. Mollenauer, M. Ho, T. Kurosaki, A. Weiss,
Mol. Cell. Biol. 27, 2732–2745 (2007).
switching rate of SOS with Ras-GTP or Ras-GDP allosterically bound is the same: gGTP = gGDP. Right
column shows simulations in which SOS with Ras-GTP allosterically bound switches more slowly than AC KNOWLED GME NTS
did SOS with Ras-GDP allosterically bound: gGTP < gGDP. For each switching condition, the Ras-GTP Major support was provided by NIH P01 AI091580. Additional
distribution is plotted for increasing cellular levels of SOS. (D) Distributions of Ras-GTP resulting from support was provided by the Danish Council for Independent
static SOS (red traces) or fluctuating SOS (blue traces). The histograms span a matrix of cellular levels Research, Natural Sciences (L.I., S.M.C., and D.S.), by the
Mechanobiology Institute, National University of Singapore
of the two RasGEFs SOS and RasGRP. Conditions in which only fluctuating SOS gives rise to a bimodal
(C.-H.Y. and J.T.G.), and by the Berkeley Education Alliance for
Ras-GTP distribution are shaded gray. Research in Singapore (J.T.G.). L.I., H.-L.T., W.-C.L., and J.T.G.
planned and designed the research; H.-L.T. and L.I. performed
GTPases by GEFs. This system is generalizable 4. J. Cherfils, M. Zeghouf, Physiol. Rev. 93, 269–309 experiments and analyzed data; S.M.C. and W.-C.L. assisted with
(2013). experiments; H.-L.T., S.M.C., and H.-J.W. developed algorithms for
to a broad class of such interactions. We antici- data analysis; H.-L.T., S.M.A., and P.T. performed stochastic
5. A. Mor, M. R. Philips, Annu. Rev. Immunol. 24, 771–800
pate that fluctuations and distributions of rates (2006). simulations; J.I., J.G., and S.D.H. prepared proteins; C.-H.Y., C.R.,
may well prove to be just as important, if not more 6. S. M. Margarit et al., Cell 112, 685–695 (2003). and R.S.P. prepared patterned substrates; L.I., H.-L.T., and J.T.G.
so, than ensemble averages in the behavior of 7. S. Boykevisch et al., Curr. Biol. 16, 2173–2179 (2006). wrote the manuscript; J.T.G. supervised the project. All
8. J. Das et al., Cell 136, 337–351 (2009). authors commented on the manuscript.
cellular signaling networks. The single-molecule
9. J. E. Jun, I. Rubio, J. P. Roose, Front. Immunol. 4, 239
membrane surface assays introduced here render (2013).
such phenomena accessible to quantitative ex- 10. H. Sondermann et al., Cell 119, 393–405 (2004). SUPPLEMENTARY MATERIALS
perimental investigation. 11. K. K. Yadav, D. Bar-Sagi, Proc. Natl. Acad. Sci. U.S.A. 107, www.sciencemag.org/content/345/6192/50/suppl/DC1
3436–3440 (2010). Materials and Methods
12. J. Gureasko et al., Proc. Natl. Acad. Sci. U.S.A. 107, 3430–3435 Figs. S1 to S18
RE FE RENCES AND N OT ES
(2010). Tables S1 to S4
1. J. T. Groves, J. Kuriyan, Nat. Struct. Mol. Biol. 17, 659–665 13. M. Tartaglia et al., Nat. Genet. 39, 75–79 (2007).
References (45–69)
(2010). 14. E. Korobkova, T. Emonet, J. M. Vilar, T. S. Shimizu, P. Cluzel,
Movies S1 and S2
2. J. D. Scott, T. Pawson, Science 326, 1220–1224 (2009). Nature 428, 574–578 (2004).
3. J. L. Bos, H. Rehmann, A. Wittinghofer, Cell 129, 865–877 15. D. J. Irvine, M. A. Purbhoo, M. Krogsgaard, M. M. Davis, Nature 3 January 2014; accepted 10 June 2014
(2007). 419, 845–849 (2002). 10.1126/science.1250373

54 4 JULY 2014 • VOL 345 ISSUE 6192 sciencemag.org SCIENCE

You might also like